Imaging Fourier-transform spectrometer measurements of a ... - AFIT

1 downloads 0 Views 925KB Size Report
Apr 15, 2014 - 2Innovative Scientific Solutions Inc., Dayton, Ohio 45459, USA. 3Air Force Research .... An interferometric “datacube” is a collection of snapshot.
2350

OPTICS LETTERS / Vol. 39, No. 8 / April 15, 2014

Imaging Fourier-transform spectrometer measurements of a turbulent nonpremixed jet flame Jacob L. Harley,1 Brent A. Rankin,2 David L. Blunck,3,5 Jay P. Gore,4 and Kevin C. Gross1,* 1

Department of Engineering Physics, Air Force Institute of Technology, Wright-Patterson AFB, Ohio 45433, USA 2 Innovative Scientific Solutions Inc., Dayton, Ohio 45459, USA 3

Air Force Research Laboratory, Aerospace Systems Directorate, Wright-Patterson AFB, Ohio 45433, USA 4

5

School of Mechanical Engineering, Purdue University, West Lafayette, Indiana 47907, USA

Currently Assistant Professor of Mechanical Engineering at Oregon State University, Corvallis, Oregon 97331, USA *Corresponding author: [email protected] Received February 17, 2014; revised March 14, 2014; accepted March 14, 2014; posted March 14, 2014 (Doc. ID 206263); published April 9, 2014 This work presents recent measurements of a CH4 ∕H2 ∕N2 turbulent nonpremixed jet flame using an imaging Fouriertransform spectrometer (IFTS). Spatially resolved (128 × 192 pixels, 0.72 mm∕pixel) mean radiance spectra were collected between 1800 cm−1 ≤ ν~ ≤ 4500 cm−1 (2.22 μm ≤ λ ≤ 5.55 μm) at moderate spectral resolution (δ~ν  16 cm−1 , δλ  20 nm) spanning the visible flame. Higher spectral-resolution measurements (δ~ν  0.25 cm−1 , δλ  0.3 nm) were also captured on a smaller window (8 × 192) at 20, 40, and 60 diameters above the jet exit and reveal the rotational fine structure associated with various vibrational transitions in CH4 , CO2 , CO, and H2 O. These new imaging measurements compare favorably with existing spectra acquired at select flame locations, demonstrating the capability of IFTS for turbulent combustion studies. © 2014 Optical Society of America OCIS codes: (120.1740) Combustion diagnostics; (110.3080) Infrared imaging; (110.4234) Multispectral and hyperspectral imaging; (300.2140) Emission; (300.6300) Spectroscopy, Fourier transforms. http://dx.doi.org/10.1364/OL.39.002350

Combustion diagnostics is a field of long standing interest with many resources continually dedicated to its study. Turbulence has significant effects on combustion processes such as turbulence-chemistry interactions, turbulence-radiation interactions, scalar dissipation, transport, and mixing. Nonintrusive optical diagnostic methods have been used to study combustion and all must consider the effects of turbulence. Laser-based methods are highly effective and widely used due to their high spectral and temporal resolution [1]. Dispersive instruments [2,3] and Fourier-transform spectrometers [4] have been used with optical scanners to tomographically deconvolve temperature and species concentrations. High-speed infrared cameras with various bandpass filters have been used to map spatial variations in radiant intensity and relate these to various measures of turbulence (e.g., integral length and time scales) [5] as well as the spatial distribution of scalar values (e.g., temperature and mole fraction) [6]. An imaging Fourier-transform spectrometer (IFTS) is a hyperspectral imager that combines a Michelson interferometer with a staring infrared focal-plane array (FPA). There are several potential advantages of this instrumentation for combustion diagnostics. High spectral resolution across a wide bandpass enables identification of multiple species. Proper interpretation of the spectrum can permit simultaneous determination of temperature and species concentrations [7]. High spectral resolution is also beneficial to tomographic reconstruction techniques [8]. High-speed broadband infrared imagery is collected during each interferometric scan. This captures turbulence information and enables similar types of analysis already performed using infrared cameras. IFTS provides a useful passive and nonintrusive technique for studying combustion and is particularly useful when (1) both high-speed imagery and spatially resolved spec0146-9592/14/082350-04$15.00/0

tra are required, (2) characterization of high-pressure systems is required and collisional broadening effects become important, (3) more than one optical port is not available, limiting the types of laser-based methods available for interrogation. The present work presents the first IFTS measurements of a canonical turbulent jet flame. The scope of this work includes a qualitative discussion of the spectral imagery and a quantitative comparison with existing spectral measurements acquired at select locations in a similar flame. The impact of turbulent intensity fluctuations on interferogram formation is also described. Quantitative interpretation of flame spectra is the ultimate goal of this effort. However, it requires scalar-field fluctuation statistics, and this important topic will be considered in future work. The experiment consisted of the Telops Hyper-Cam IFTS, two calibration blackbodies, and the flame. The flame tube is 480 mm long with an 8 mm exit diameter (D), mounted vertically, and moveable via unislide to allow combined imaging of the entire visible flame length without camera tilt. The flame replicates Flame DLR_A from the International Workshop on Measurement and Computation of Turbulent Nonpremixed Flames (TNF Workshop) with a jet exit Reynolds number of 15,200 and exit velocity of 42.2 m∕s. Mass flow rates were 313, 59, 1105 mg∕s for CH4 , H2 , and N2 , respectively. Flow rates were calibrated using a dry turbine meter and controlled by setting the pressure upstream of three choked orifice plates [5]. The TNF Workshop flames are well characterized and designed for collaborative comparisons of measurements and models. A library of local velocities and scalar values (temperature, species mole fractions) measured simultaneously using laser Doppler velocimetry, Raman, Rayleigh, and LIF techniques is available for download [9,10]. © 2014 Optical Society of America

April 15, 2014 / Vol. 39, No. 8 / OPTICS LETTERS

The IFTS is based on a traditional Michelson interferometer coupled to a high-speed 320 × 256 indium antimonide staring FPA via f ∕#  2.5 imaging optics [7,11]. The spectral range covers 1800–6667 cm−1 , and the spectral resolution can be selected between 0.25 and 150 cm−1 . An interferometric “datacube” is a collection of snapshot images taken at equally spaced optical path differences (OPDs), and Fourier-transformation along this dimension produces a spectrum at each pixel. An external 0.25× telescope expanded the field-of-view and reduced the minimum working distance to the flame. A 45% transmission neutral density filter, used to prevent saturation, limited the short-wavelength response to 2.22 μm (4500 cm−1 ). The IFTS was located 47.5  1.0 cm from the flame. The imaging system has an effective focal length of 19.7 mm at this working distance. The 30 μm pixel pitch of the FPA yields an instantaneous field-of-view (IFOV) of 1.52 mrad which translates to 0.72  0.02 mm at the flame and is constant across the array. The mean RMS spot size radius is 13.7 μm, and increases from 11.2 to 21.1 μm moving from center to corner of a 128 × 192 window. Mapping the Rayleigh λ∕4 wavefront error depth-of-focus criterion, δf  2λ f ∕#2 , to object space produces a conservative estimate of the depth-of-field of 2 cm when computed at 2.5 μm, the shortest wavelength with appreciable energy arriving at the FPA. Throughout much of the flame, the spectral imagery can thus be interpreted as integrated along the line-of-sight (LOS). However, the widest part of the flame is ∼15 cm, indicating some blurring will occur along the LOS. A detailed Zemax [12] optical model of our system indicates that more than 75% of the energy (relative to the diffraction-limited case, 86.4%) comes from the LOS for a pixel viewing the center of the widest (7.5 cm) flame region. The IFTS was mounted to a gimbal with preset locations for intermittent calibration measurements. A standard two-point calibration using the wide-area blackbodies set to 595°C and 200°C was performed pixel-wise to determine the system response [gain, Gi ~ν] and instrument self-emission [offset, LIi ~ν]. The higher blackbody temperature produced a peak signal at ∼90% of the detector’s dynamic range and slightly exceeding that from the bright-

2351

est part of the flame. At 595°C, the Planckian distribution monotonically decreases with frequency across the detector bandpass. This resulted in a nominal signal-to-noise ratio (SNR) in G~ν, which decreased nearly linearly from 15 to 1 between 3000 and 5000 cm−1 . Since the system response is known to vary smoothly and slowly with ν~ , a spline was fit to each pixel’s gain curve to mitigate the impact of low-gain SNR on the calibrated spectrum. Two sets of flame measurements were made. The first set was collected with high spectral resolution (0.25 cm−1 ) in a small window (8 × 192) traversing the flame at 20, 40, and 60D above the burner to facilitate identification of various chemical species. Interferometric datacubes consisted of 52,742 images and were collected at a rate of 0.55 Hz, and 512 cubes were averaged to produce a mean, calibrated image of the flame radiance. The second set increased the FPA window height (128 × 192) to facilitate measurement of the entire flame and decreased spectral resolution (16 cm−1 ) to simplify data reduction. Datacubes consisted of 1186 images and the acquisition rate increased to 4.2 Hz. Seven separate regions of the flame were imaged to produce a composite image of the entire flame. In each set the camera’s integration time was 20 μs, and imaging frame rates exceeded 5 kHz. Ambient temperature, pressure, and humidity were monitored with a Kestrel 4500 Weather Meter with averages of 25°C, 989 hPa, and 44% respectively. Fourier-transform spectrometry is typically used to study static scenes, so we briefly review interferogram formation so that the impact of stochastic intensity variations from the turbulent flame can be properly understood. The formation of an interferometric datacube is depicted in Panel A of Fig. 1. Light enters the Michelson producing an interference pattern at the FPA which encodes the spectral radiance at each pixel in the image. This interferogram is a function of the OPD, x, or time, t (the two being related by the constant mirror sweep speed v): Z I i x 

∞ 0

1  cos2π ν~ xGi ~νLSi ~ν  LIi ~νd~ν; (1)

AC  I DC i  I i x:

(2)

Fig. 1. Panel A: schematic illustrating the FPA capturing infrared images at fixed OPDs as the Michelson sweeps, generating an interferometric datacube. Panel B: single interferogram (green, upper curve) and corresponding raw spectrum (red, lower curve) at flame center 20D above exit. Panel C: Time-averaged interferogram (green, upper curve) and corresponding mean flame spectrum (red, lower curve).

2352

OPTICS LETTERS / Vol. 39, No. 8 / April 15, 2014

Here, LSi ~ν is the scene spectrum at pixel i, LIi ~ν represents the instrument’s thermal self-emission, and Gi ~ν is the system response. A two-point calibration determines term represents the spectrally Gi ~ν and LIi ~ν. The I DC i integrated intensity and I AC i x is the cosine transform of the spectrum produced by the Michelson. In a nonimaging FTS, the detector is often AC coupled, dedicating the full range of the analogue-to-digital converter to the more useful AC piece. This is not possible with an FPA, so each pixel has a modulation signal riding on top of the DC offset. For a turbulent jet the scene radiance is stochastically fluctuating on a timescale much shorter than the interferometer’s acquisition rate. Thus, the “DC” term is now time-dependent and the AC term has no simple interpretation as it is the cosine transformation of a stochastically varying signal. This is illustrated in Panel B of Fig. 1 showing a single interferogram and corresponding raw magnitude spectrum. The fluctuations in integrated intensity dominate the signal and obscure the zero-path difference (ZPD) where all wavelengths constructively interfere. The corresponding spectrum is dominated by the frequencies associated with turbulent radiation fluctuation, although a feature near 2300 cm−1 resembling emission from the asymmetric stretching mode (ν3 ) of CO2 is recognizable. Large intensities below the detector cut-off (~ν ≤ 1800 cm−1 ) are due to turbulent fluctuations. For an ergodic system, an ensemble of measurements will produce a mean interferogram corresponding to the mean spectral radiance since Eq. (1) is a linear transformation. Panel C presents the same pixel’s mean interferogram from 512 measurements, demonstrating that the turbulent fluctuations are suppressed. The resulting spectrum is now recognizable with rotational fine structure associated with vibrational transitions in H2 O, CH4 , CO, and CO2 . Figure 2 presents uncalibrated broadband imagery in Panels A and B, dividing the flame along the axis of symmetry into single-snapshot and time-averaged quantities. Each segment is temporally independent from the others. In Panels A and B the images were acquired at a common

OPD near x  370 μm. Away from ZPD the imagery is similar to what an infrared camera would measure (I i x ≃ I DC i ) since the broadband nature of radiation enDC sures jI AC i xj ≪ I i . At the burner tip, the distance traveled by the jet during the FPA’s integration time is 0.84 mm, exceeding the IFOV by approximately 12%, a conservative estimate of blurring due to the rapid deceleration of the jet. Moreover, the turbulence integral length scales for this flame between 20 and 60D are within 9.1–24 mm [5]. Thus, the turbulent structures exceed the spatial resolution by an order of magnitude. The time between repeated observations at a particular OPD is 240 ms, greatly exceeding the turbulence integral time scales (2.3–5 ms between 20 and 60D). Repeated observations at each OPD are statistically independent. Also shown in Fig. 2 are calibrated images (via the time-averaged interferometric cubes) resulting from integration over spectral bands corresponding to CO2 (Panel C) and CH4 (Panel D). The imagery provides a qualitative impression of the distribution of these species throughout the flame. For example, CO2 emission is low near the burner tip but increases axially as air is entrained to fuel the combustion reaction. The CO2 band radiance peaks at 55D, in agreement with previous narrow-band 4.34  0.10 μm measurements of the same flame [5]. Temperature, species concentration and the LOS distance through the flame all affect the band-integrated radiance. For example, the initial increase in CH4 intensity with distance from the tip is due to increasing temperature despite decreasing concentration. Proper interpretation of the spectrum will enable the deconvolution of these interdependencies. The mean high-resolution spectrum acquired from a diametric path at 20D is presented in Panel A of Fig. 3. Emissions from CH4 , CO, CO2 , and H2 O are resolved, and major vibrational transitions are annotated using common spectroscopic notation [13]. Individual lines associated with the P-branch of CO are visible between 2000 and 2150 cm−1 ; lines from the R-branch overlap with the strong CO2 emission band associated with the asymmetric stretching mode. The mean low-resolution spectra

Fig. 2. Panel A: single broadband images from the lower spectral resolution (16 cm−1 ) datasets with 128 × 192 FPA window. Panel B: corresponding time-averaged broadband images. Panel C: Time-averaged radiance spectrally averaged over a prominent CO2 band. Panel D: Time-averaged radiance spectrally averaged over a prominent CH4 band. (Last spatial region limited due to unislide range.)

April 15, 2014 / Vol. 39, No. 8 / OPTICS LETTERS

Fig. 3. Panel A: diametric, high-resolution (δ~ν  0.25 cm−1 ) apparent flame spectrum at 20D with spectroscopic transitions annotated. Panel B: apparent (—) and atmospheric-corrected (  ) low-resolution diametric flame spectra (δ~ν  16 cm−1 ) at 20, 40, and 60D (black, red, blue) compared with previous measurements (○). Radiance uncertainty (95% confidence interval) presented as a translucent band around each apparent spectrum. The CO2 and CH4 bands used in Fig. 2 are identified.

from diametric paths at 20, 40, and 60D are compared with previous (nonimaged) LOS measurements of flame A by Zheng et al. [2] in Panel B of Fig. 3. A spatial average over a 3 × 3 window was performed to approximate the 2 mm resolution of Zheng’s data. The solid lines represent the apparent (i.e., at-sensor) radiance. For proper comparison, atmospheric correction was performed (dashed lines) using measured lab conditions and assuming 500 ppm CO2 concentration. Agreement is excellent at 40 and 60D. However, the CO2 peak near 2300 cm−1 is ∼20% below the previously reported value at 20D. Radiance uncertainties (95% confidence interval) are shown and include the effects of both systematic errors in calibration and noise. Noise is estimated as the rootmean-square value of the imaginary component of the time-averaged spectrum. Turbulent fluctuations are minimized in the time-averaged interferogram, so the error band does not quantify the large variance in flame radiance. At 20 and 60D, the mean uncertainty between 2200

2353

and 2350 cm−1 is 5% and 4%, respectively. Between 3000 and 4000 cm−1 , the uncertainty increases to 20% and 10%, respectively, at 20 and 60D. This work makes two key contributions. First, it validates the use of IFTS for studying turbulent flames and provides complete time-averaged hyperspectral imagery of flame A. Segments of the flame were imaged with spectral resolution an order-of-magnitude better than previous measurements, and this data could be used to evaluate and improve narrowband radiation models. Second, it demonstrates the potential of IFTS for combustion diagnostics. Mean hyperspectral images contain information about the distribution of both temperature and many major gas species throughout the flame. Additionally, the high-speed broadband imagery comprising each interferometric measurement contains information about the fluctuation statistics. High resolution in all three domains—spectral, spatial, and temporal—is extremely valuable in the study of turbulent combustion and is captured in IFTS measurements. Having demonstrated the validity of time-averaged spectra in this work, our efforts will now turn to leveraging the high-speed imagery contained within IFTS measurements to understand scalar fluctuation statistics, a key step in the quantitative interpretation of turbulent flame hyperspectral imagery. References 1. K. Kohse-Höinghaus and J. B. Jeffries, Applied Combustion Diagnostics (Taylor & Francis, 2002). 2. Y. Zheng, R. S. Barlow, and J. P. Gore, J. Heat Transfer 125, 678 (2003). 3. Y. Zheng, R. S. Barlow, and J. P. Gore, J. Heat Transfer 125, 1065 (2003). 4. P. R. Solomon, P. E. Best, R. M. Carangelo, J. R. Markham, P.-L. Chien, R. J. Santoro, and H. G. Semerjian, Symp. Int. Combust. Proc. 21, 1763 (1988). 5. B. A. Rankin, D. A. Blunck, and J. P. Gore, J. Heat Transfer 135, 021201 (2013). 6. D. Blunck, S. Basu, Y. Zheng, V. Katta, and J. Gore, Proc. Comb. Inst. 32, 2527 (2009). 7. K. C. Gross, K. C. Bradley, and G. P. Perram, Environ. Sci. Technol. 44, 9390 (2010). 8. L. Ma, W. Cai, A. W. Caswell, T. Kraetschmer, S. T. Sanders, S. Roy, and J. R. Gord, Opt. Express 17, 8602 (2009). 9. http://www.sandia.gov/TNF/abstract.html. 10. W. Meier, R. S. Barlow, Y.-L. Chen, and J.-Y. Chen, Combust. Flame 123, 326 (2000). 11. V. Farley, A. Vallières, M. Chamberland, A. Villemaire, and J.-F. Legault, Proc. SPIE 6398, 63980T (2006). 12. http://radiantzemax.com. 13. J. M. Hollas, High Resolution Spectroscopy, 2nd ed. (Wiley, 1998).